You are on page 1of 8

The exact molecular wavefunction as a product of an electronic and a nuclear

wavefunction
Lorenz S. Cederbaum

Citation: The Journal of Chemical Physics 138, 224110 (2013); doi: 10.1063/1.4807115
View online: http://dx.doi.org/10.1063/1.4807115
View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/138/22?ver=pdfcov
Published by the AIP Publishing

Articles you may be interested in
Response to Comment on Correlated electron-nuclear dynamics: Exact factorization of the molecular
wavefunction' [J. Chem. Phys.139, 087101 (2013)]
J. Chem. Phys. 139, 087102 (2013); 10.1063/1.4818523

Molecular structure calculations: A unified quantum mechanical description of electrons and nuclei using
explicitly correlated Gaussian functions and the global vector representation
J. Chem. Phys. 137, 024104 (2012); 10.1063/1.4731696

Electronic excited-state energies from a linear response theory based on the ground-state two-electron reduced
density matrix
J. Chem. Phys. 128, 114109 (2008); 10.1063/1.2890961

Cumulant reconstruction of the three-electron reduced density matrix in the anti-Hermitian contracted
Schrdinger equation
J. Chem. Phys. 127, 104104 (2007); 10.1063/1.2768354

A computationally efficient exact pseudopotential method. II. Application to the molecular pseudopotential of an
excess electron interacting with tetrahydrofuran (THF)
J. Chem. Phys. 125, 074103 (2006); 10.1063/1.2218835


This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
180.149.51.67 On: Mon, 19 May 2014 08:56:20
THE JOURNAL OF CHEMICAL PHYSICS 138, 224110 (2013)
The exact molecular wavefunction as a product of an electronic
and a nuclear wavefunction
Lorenz S. Cederbaum
a)
Theoretische Chemie, Physikalisch-Chemisches Institut, Universitt Heidelberg, Im Neuenheimer Feld 229,
D-69120 Heidelberg, Germany
(Received 10 January 2013; accepted 6 May 2013; published online 13 June 2013)
The Born-Oppenheimer approximation is a basic approximation in molecular science. In this ap-
proximation, the total molecular wavefunction is written as a product of an electronic and a nuclear
wavefunction. Hunter [Int. J. Quantum Chem. 9, 237 (1975)] has argued that the exact total wave-
function can also be factorized as such a product. In the present work, a variational principle is
introduced which shows explicitly that the total wavefunction can be exactly written as such a prod-
uct. To this end, a different electronic Hamiltonian has to be dened. The Schrdinger equation for
the electronic wavefunction follows from the variational ansatz and is presented. As in the Born-
Oppenheimer approximation, the nuclear motion is shown to proceed in a potential which is the
electronic energy. In contrast to the Born-Oppenheimer approximation, the separation of the center
of mass can be carried out exactly. The electronic Hamiltonian and the equation of motion of the
nuclei resulting after the exact separation of the center of mass motion are explicitly given. A simple
exactly solvable model is used to illustrate some aspects of the theory. 2013 AIP Publishing LLC.
[http://dx.doi.org/10.1063/1.4807115]
I. INTRODUCTION
The Born-Oppenheimer approximation
1, 2
is a milestone
in the theory of molecules and of electronic matter in gen-
eral. The much larger masses of the nuclei compared to that
of electrons allow for an approximate separation of the elec-
tronic and nuclear motions and this separation simplies the
quantum as well as classical treatment of molecules substan-
tially. One should be aware that even the notion of molecular
electronic states is connected to this approximation.
As usual, one writes the molecular Hamiltonian
H = T
N
+H
el
(1)
as a sum of the kinetic energy operator T
N
of the nuclei
and the electronic Hamiltonian H
el
which governs the elec-
tronic motion at xed nuclear congurations. Molecules are
rather complicated quantum objects and the solution of the
Schrdinger equation for the total molecular Hamiltonian H
in (1) is, in general, rather involved. In the Born-Oppenheimer
approximation, the ansatz

BO
= (r; R)(R) (2)
for the wavefunction is a product of an electronic wave-
function and a nuclear wavefunction . Here, r and R
stand symbolically for all electronic and nuclear coordinates,
respectively. The electronic Schrdinger equation
H
el
= E
el
(R) (3a)
denes an electronic energy E
el
(R) which, of course, depends
on the nuclear coordinates. The electronic wavefunction
a)
Electronic mail: lorenz.cederbaum@pci.uni-heidelberg.de
(r; R) is normalized at all values of the nuclear coordinates
_
||
2
dr = |
el
= 1. (3b)
The index el indicates that the integration is over the elec-
tronic coordinates r only.
The main idea behind the Born-Oppenheimer approxima-
tion is that due to the smallness of the mass of the electrons
compared to that of the nuclei, the electronic wavefunction
varies very smoothly with R and is not affected by the oper-
ation of T
N
. If so, one can insert the ansatz (2) into the full
Schrdinger equation, integrate out the electronic contribu-
tion and obtain the illuminating result
[T
N
+E
el
(R)] = E
BO
, (4)
where E
BO
is the total energy of the molecule in the Born-
Oppenheimer approximation. The essence of this central
equation is that the nuclei move in the potential provided by
the electronic energy E
el
(R). This quantity is hence called the
potential energy surface. In their recent paper, Sutcliffe and
Woolley
3
discuss that this surface does not arise naturally
from the solution of the Schrdinger equation for the molecu-
lar Coulomb Hamiltonian and present arguments substantiat-
ing their view.
The Born-Oppenheimer approximation as provided by
Eqs. (2)(4) has been extremely useful in numerous applica-
tions and is generally widely applied. It has become a standard
reference even in cases where it fails. This approximation
does fail, often severely, in particular for polyatomic
molecules whenever the potential energy surfaces belong-
ing to different electronic states come close to each other.
The most dramatic failure is encountered when these sur-
faces exhibit conical intersections.
4
The nuclear motion then
0021-9606/2013/138(22)/224110/7/$30.00 2013 AIP Publishing LLC 138, 224110-1
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
180.149.51.67 On: Mon, 19 May 2014 08:56:20
224110-2 Lorenz S. Cederbaum J. Chem. Phys. 138, 224110 (2013)
proceeds over the potential surfaces of the involved elec-
tronic states which are now coupled by this motion. To de-
scribe the nuclear motion, one usually introduces a group
Born-Oppenheimer approximation for the manifold of cou-
pled electronic states.
48
The total wavefunction is then a sum
of products of electronic and nuclear wavefunctions.
Can the Born-Oppenheimer ansatz (2) be improved with-
out resorting to a sum of products of electronic and nu-
clear wavefunctions? The idea to search for an improved total
wavefunction which is a single product as in (2) is rather old.
Pack and Hirschfelder
9
searched in 1970 for the best Born-
Oppenheimer-like approximation. Hunter
10
argued that the
exact wavefunction can be factorized as in (2)
= (r; R)f (R) (5a)
dening a nuclear wavefunction f (R) via the so-called
marginal probability
|f (R)|
2
= |
el
. (5b)
An electronic wavefunction can be constructed using
= /f (5c)
provided that f (R) is nodeless and the normalization condition
|
el
= 1 (5d)
is chosen. Then, the electronic wavefunction is well dened
and a potential energy surface for f (R) could be dened as
well as
|H|
el
. (5e)
The nodeless function f (R) has been further analyzed by
Hunter
11
and several numerical studies on the potential in
Eq. (5e) have been performed for the H
2
molecule and its
ion,
12, 13
see also Ref. 14 on a related potential. The situation
is nicely described in a recent general paper by Sutcliffe.
15
The total wavefunction must be known already before-
hand in order to construct by the above procedure the elec-
tronic and nuclear wavefunctions and f as well as the po-
tential in Eq. (5e). That is, these quantities are constructed
a posteriori after the full solution of the Schrdinger equation
has been determined.
In the present work, we would like to determine the elec-
tronic and nuclear wavefunctions entering the product ansatz
of the exact total wavefunction from rst principles. To that
end, we introduce a variational principle which leads to the
exact factorized result and provides the Schrdinger equations
for the electronic and nuclear wavefunctions. The electronic
wavefunction will, of course, not be the eigenfunction of the
electronic Hamiltonian H
el
we are used to. For that purpose,
we will have to introduce a different, more involved electronic
Hamiltonian. We shall show that the nuclear motion in that
case still proceeds over a potential surface provided by the
electronic energy. The basic ideas are collected in Sec. II. In
Sec. III, we apply them to a simple exactly solvable model
16
which has been used in the literature to study the ingredients
of the Born-Oppenheimer approximation. An application to a
more realistic problem has been performed but is beyond the
scope of the present paper.
Finally, we would like to mention that the Born-
Oppenheimer approximation is also widely used in the con-
text of nuclear dynamics. In this case, the time-dependent
Schrdinger equation is solved for the nuclear motion in the
potential surface E
el
(R). In recent years, theoretical interest
in electronic dynamics has emerged as well,
1729
rst driven
by curiosity and later on motivated by the advent of sub-
femtosecond and attosecond experimental technologies.
3035
To incorporate both the nuclear and the electronic dynam-
ics, very recently the full time-dependent Born-Oppenheimer
approximation has been introduced and investigated.
36
Here,
both the electronic and nuclear wavefunctions depend on
time. To go beyond this approximation, a time-dependent
group Born-Oppenheimer approximation
36
and an exact fac-
torization of the wavepacket into an electronic and a nuclear
wavepackets
37, 38
have been put forward.
II. BASIC IDEAS
A. The factorization
We start with the total Hamiltonian (1) and its eigenvalue
equation
H = E. (6)
The basic idea is to exactly factorize into an electronic
wavefunction (r; R) and a nuclear wavefunction (R)
= (r; R)(R), (7a)
where the new electronic function is normalized
|
el
= 1 (7b)
similar to the eigenfunction of H
el
in Eqs. (3a) and (3b).
Choosing the total wavefunction to be normalized, i.e.,
| = 1, the nuclear function is normalized in nuclear
space. To proceed, we have to specify the kinetic energy op-
erator T
N
. For transparency, we use conveniently scaled rect-
angular nuclear coordinates, for which T
N
can be expressed
as
T
N
=

2
2M
=

2
2M
, (8)
where is the gradient in nuclear space, the dot is the com-
mon scalar product, and M is an average nuclear mass.
Inserting the ansatz (7) into the Schrdinger equation (6)
and using the above expression for the kinetic energy oper-
ator, we obtain a lengthy equation coupling and . This
equation simplies considerably if we introduce an electronic
operator H
el
which has as its eigenstate
H
el
= E
el
(R). (9)
This is readily accomplished by choosing
H
el
= H
el


2
2M
(ln ) +T
N
. (10)
The electronic energy E
el
(R) can be easily found by taking
the expectation value |H
el
|
el
of the effective electronic
Hamiltonian H
el
with the electronic state . We do not con-
ne ourselves to real solutions of Eq. (10), i.e., can be com-
plex. But, for real functions , the term ||
el
vanishes
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
180.149.51.67 On: Mon, 19 May 2014 08:56:20
224110-3 Lorenz S. Cederbaum J. Chem. Phys. 138, 224110 (2013)
because of the normalization condition (7b), and hence
E
el
(R) = |H|
el
, (11)
i.e., the electronic energy is just the electronic expectation
value of the full Hamiltonian H taken with the new electronic
function .
Making use of Eq. (9), the Schrdinger equation (6) with
= boils down exactly to
[T
N
+E
el
(R)] = E. (12)
The operator T
N
+E
el
(R) governs the nuclear motion in the
state and provides the exact energy E of the system. As
in the Born-Oppenheimer approximation, Eq. (4), the nuclear
motion proceeds in a potential which is provided by the elec-
tronic energy.
At this point, a number of issues have to be discussed.
Some authors call Eq. (4) the Born-Oppenheimer adiabatic
approximation
7
and reserve the name Born-Oppenheimer ap-
proximation to an equation as (4), but with a diagonal cor-
rection term = |T
N
|
el
added to the electronic energy
E
el
= |H
el
|
el
. One notices that E
el
+ is nothing but
the expectation value |H|
el
of the full Hamiltonian taken
with the usual electronic wavefunction . The new electronic
energy E
el
in (11) is nothing but the analogous quantity com-
puted, however, with the exact electronic wavefunction . See
also Eq. (5e).
Although is an electronic wavefunction which depends
on R, it is not true in the conventional sense that this is a pure
parametrical dependence, since the nuclear momentum oper-
ators appear in the electronic Hamiltonian (10). This implies
that the electronic energy in Eq. (10) which serves as the po-
tential surface for the nuclear motion in Eq. (11) implicitly
already contains some information on the nuclear motion. For
general molecules, it will be quite challenging to solve the
problem, since the product ansatz basically undoes the Born
Oppenheimer separation.
An important issue is the appearance of ln in the elec-
tronic Hamiltonian H
el
. For the ground state of the nuclear
motion, the nuclear wavefunction can be expected to be node-
less. Then, ln is well dened everywhere. For other states,
may have nodes, and ln will have singularities. Apart from
these singularities, Eq. (10) for H
el
is well dened. This equa-
tion can be used to determine everywhere except at these
singularities. This should be sufcient to solve the problem.
In Subsection II B, we shall derive the working equations also
from another perspective, namely, by employing variational
theory. It will be shown that the singularities are suppressed
by a multiplicative term
2
.
Another aspect of the appearance of ln in H
el
is the
fact that the electronic wavefunction in the exact factorization
= depends on . Consequently, Eq. (9) for the elec-
tronic wavefunction and Eq. (12) for the nuclear wavefunc-
tion are intimately coupled. This lowers the practicability of
the ansatz in general. Nevertheless, the ansatz provides addi-
tional insight into the complex problem of understanding the
molecular Schrdinger equation. It could also be that the ap-
proach is amenable to further simplications, exact or inexact
but accurate. Here, we would just like to mention that as
depends on , the electronic energy is a functional of , i.e.,
E
el
= E
el
[]. In turn, the appealing equation (12) for the nu-
clear motion can also be seen as
{T
N
+E
el
[]} = E. (13)
If one nds a scheme to determine the functional form of
E
el
[], this equation could become of practical relevance.
B. Variational approach
In this subsection, we start from the Schrdinger
equation (6) for the exact wavefunction and attempt to solve it
variationally with the product ansatz (7). Since depends on
r and parametrically on R, which is expressed by the nor-
malization condition (7b) valid at any value of R, a variational
ansatz is not trivial.
We introduce the following functional:
[, ] = |H| +(1 |)
+(1 |
R
), (14)
where and are Lagrange parameters ensuring that |
= 1 and
_
||
2
dR = 1. If we want to vary |H| with re-
spect to (r; R), we cannot use the constraint |
el
= 1 be-
cause this constraint is only over r and we have to vary over
r and R. With |
el
= 1 the constraints in (14) are easy to
fulll. On the other hand, |
el
= 1 does not follow imme-
diately from those in (14) because depends not only on r,
but also on R.
Varying with respect to

, i.e.,

= 0, (15a)
immediately gives

(H ) = 0. (15b)
Remembering that H = H
el
+ T
N
and applying T
N
on
leads to
||
2
_
H
el


2
M
(ln ) +T
N

_
= (

T
N
),
(15c)
where (

T
N
) implies that T
N
is applied to only and
does not act on . In parenthesis, we re-discover the effective
electronic Hamiltonian H
el
(10).
To proceed, we now vary [, ] with respect to

and
set

= 0. (16a)
This readily leads to
|H|
el
= [ +|
el
]. (16b)
By applying again T
N
on , we can rewrite this equation to
give
|
el
_
T
N
+
|H
el
|
el
|
el
_
= [ +|
el
]. (16c)
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
180.149.51.67 On: Mon, 19 May 2014 08:56:20
224110-4 Lorenz S. Cederbaum J. Chem. Phys. 138, 224110 (2013)
Multiplying from the left by

allows us to obtain
(

T
N
) =
||
2
|
el
_
+|
el

|H
el
|
el
|
el
_
,
(16d)
which can be inserted into the right hand side of Eq. (15c) and
now this equation takes on the appearance
||
2
H
el
=
||
2
|
el
_
|H
el
|
el
|
el

_
. (17a)
This equation can be used to obtain by multiplying it from
the left by

and integrating over the electronic coordinates


||
2
= ||
2
_
|H
el
|
el
|
el
|H
el
|
el
_
. (17b)
Since should be a number and all other quantities appearing
in (17b) are functions of R, the only solution is
|
el
= 1
and hence = 0.
We are now in the position to write down the nal results
of the variational computation. Equation (17a) now reads
||
2
[H
el
E
el
(R)] = 0, (18a)
which is identical to Eq. (9) derived in Subsection II A except
for the overall factor ||
2
in front of the equation. This factor
suppresses the possible singularities in H
el
. Equation (16c)
now takes on the appearance
[T
N
+E
el
(R)] = . (18b)
This central equation which governs the nuclear motion in
the electronic state is identical to Eq. (12) obtained in
Subsection II A. From the comparison, it is evident that
= E, the exact total energy of the system.
Note added: After this work was submitted, it was found
that there is an unpublished work that provides and discusses
the equation of motion determining the exact electronic wave-
function before the separation of the center of mass.
47
C. Center of mass motion and separation
of coordinates
It is well known that in the absence of external forces,
the center of mass motion of the whole system separates
from the internal motion. In the Born-Oppenheimer approx-
imation, this separation is violated. Here, only the center of
mass of the nuclei separates. In an external magnetic eld,
the center of mass does not separate
39, 40
and the ensuing cou-
pling between this motion and the internal motion gives rise to
dramatic effects in strong elds.
4143
In magnetic elds,
the Born-Oppenheimer approximation severely fails and has
to be corrected.
4446
Since the present approach of factorizing the total wave-
function into a product of an electronic and a nuclear wave-
function is formally exact, the separation of the center of mass
is exact too. The explicit example presented in Sec. III illus-
trates this nicely. To exploit this separation, we express T
N
as the sum of the translational kinetic energy and the kinetic
energy T
n
of the internal nuclear motion
T
N
= T
n
+
1
2M
NT
_

k
P
k
_
2
, (19a)
where M
NT
is the total mass of the nuclei and P
k
is the mo-
mentum operator of the kth nucleus. Let the total momentum
of the system be P
CM
. The translational kinetic energy then
reads
1
2M
NT
_
P
CM

i
p
i
_
2
, (19b)
where p
i
is the momentum of the ith electron. Because of the
separability of the center of mass motion in eld free space,
P
CM
= K where K is the good quantum number of the
translational motion. This helps to remove exactly all the cen-
ter of mass operators from the calculations.
We shall start from our basic equations (9) and (12) and
seek for wavefunctions
K
and
K
which satisfy these equa-
tions and lead to the expected motion of the center of mass
of the system. Since
K
is an electronic and
K
a nuclear
wavefunction, we set

K
= e
i K
m
M
T

i
(r
i
R
N
)
, (20a)

K
= e
i K R
N
, (20b)
where m is the electron mass and M
T
is the total mass of the
system, i.e., the nuclei plus the electrons. The coordinate of
the ith electron is r
i
and R
N
is the center of mass of the nuclei
alone. Obviously, the total wavefunction
K
=
K

K
fullls
the known relationship

K
= e
i K R
CM
. (20c)
Here, R
CM
is the center of mass of the whole systemincluding
the electrons.
Inserting the ansatz (20a) into Eq. (9) for the electronic
wavefunction, one can eliminate the exponential function. For
that purpose, one has also to let the electronic kinetic energy
operator in H
el
operate on this exponential function and to
evaluate the term (ln
K
) appearing in the electronic
Hamiltonian H
el
in (10) using (20b). Finally, one has also
to compute the impact of the nuclear kinetic energy T
N
on the
exponential function. After a somewhat lengthy calculation,
we obtain
H
el

K
= e
i K
m
M
T

i
(r
i
R
N
)
_
H
el


2
M
(
n
ln )
n
+

M
T
K P
CM


2
K
2
2M
T
Nm
M
NT
_

and since does not depend on the center of mass of the


whole system, P
CM
= 0. The nal equation for now reads
again
H
el
= E
K
el
, (21a)
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
180.149.51.67 On: Mon, 19 May 2014 08:56:20
224110-5 Lorenz S. Cederbaum J. Chem. Phys. 138, 224110 (2013)
where the electronic Hamiltonian H
el
now takes on the ap-
pearance
H
el
= H
el


2
M
(
n
ln )
n
+T
N
+C
K
, (21b)
which, using Eqs. (19a) and (19b), can be further reduced to
give
H
el
=H
el


2
M
(
n
ln )
n
+T
n
+
1
2M
NT
_

i
p
i
_
2
+C
K
.
(21c)
The only dependence on the total translational momentumK
is via the constant
C
K
=

2
K
2
2M
T
Nm
M
NT
, (21d)
where N is the number of electrons. The quantity Nm/M
NT
is
thus the ratio of the total electron mass to that of the nuclei.
The electronic energy becomes
E
K
el
(R) = E
el
(R) +C
K
. (21e)
The electronic energy E
el
fullls as before (11) for
real .
We now turn our attention to the nuclear wavefunction

K
. Inserting the ansatz (20b) into the basic equation (12) and
recognizing that the translational kinetic energy of the nuclei
can be expressed as


2
2M
NT

2
R
2
N
,
one readily arrives at
_
T
n
+E
K
el
(R) +

2
K
2
2M
NT
_
= E, (22a)
which leads via (21e) and (21d) to the nal result
_
T
n
+E
el
(R) +

2
K
2
2M
T
_
= E. (22b)
Equations (21)(22b) complete our separation of the cen-
ter of mass motion. In these equations, only the internal nu-
clear degrees of freedom appear. Equations (22a) and (22b)
are as one would expect: The exact energies are the sum of
the internal energy and the total center of mass energy of the
system. The product ansatz = holds also in the pres-
ence of external elds. The pseudo-separation of the center of
mass motion is expected, however, to lead to intricate work-
ing equations reecting the residual interaction between the
internal and center of mass motions.
III. A SIMPLE EXACTLY SOLVABLE MODEL
In order to examine the quality of the Born-Oppenheimer
approximation, Moshinsky and Kittel (MK)
16
devised a sim-
ple model of a light particle (the electron) and two heavy parti-
cles (the nuclei) which are coupled to each other by harmonic
forces. The MK-model can be solved exactly and also explic-
itly in the Born-Oppenheimer approximation. The Hamilto-
nian of the system reads
H =
p
2
2m
+
P
2
1
2M
+
P
2
2
2M
+
k
2
(r R
1
)
2
+
k
2
(r R
2
)
2
+

2
(R
1
R
2
)
2
, (23)
where p is the momentum of the light particle of mass m, P
1
and P
2
the momenta of the heavy particles of mass M, and k
and are spring constants. The meaning of the coordinates r,
R
1
, and R
2
is obvious.
As usual, one denes the electronic Hamiltonian
H
el
=
p
2
2m
+V
el
(r),
(24)
V
el
(r) =
k
2
(r R
1
)
2
+
k
2
(r R
2
)
2
+

2
(R
1
R
2
)
2
,
which depends parametrically on the nuclear positions R
1
and
R
2
. The electronic potential is easily rewritten to give
V
el
(r) =
k
2
(r R
N
)
2
+
2+k
4
(R
1
R
2
)
2
, (25)
where R
N
= (R
1
+ R
2
)/2 is the center of mass of the nu-
clei. Since the R
i
, i = 1, 2, are parameters, the electronic
Schrdinger equation
H
el
= E
el
(| R
1
R
2
|) (26a)
is readily solved for all electronic states. The solutions
q
(r
R
N
) are harmonic oscillator eigenstates with energies
E
el,q
(| R
1
R
1
|) =
el
(q + 3/2) +
2+k
4
(R
1
R
2
)
2
,
(26b)

el
= [2k/m]
1/2
,
where, as usual, q = 0, 1, 2, . . . .
In the next step of the Born-Oppenheimer approach, one
solves Eq. (4) for the nuclear function, which in the MK
model reads
_
P
2
1
2M
+
P
2
2
2M
+E
el,q
(| R
1
R
2
|)
_
= E
BO
. (27a)
The nuclear kinetic energy operator can be written as
T
N
=
P
2
1
2M
+
P
2
2
2M
=

2
n
M
+

2
N
4M
, (27b)
where
n
= ( P
1
P
2
)/2 is the momentum conjugated to
the relative coordinate R
1
R
2
describing the internal nu-
clear motion, and
N
= P
1
+ P
2
is conjugated to the center
of mass coordinate R
N
= (R
1
+ R
2
)/2. By introducing these
momenta, these two motions are separated and the solutions
of Eq. (27a)

q,,K
(| R
1
R
2
|, R
N
) = e
i K R
N

q,
(| R
1
R
2
|) (27c)
are plane waves which describe the translational motion of
the center of mass of the nuclei and the three-dimensional
harmonic oscillator functions describing the vibrational and
rotational motion of the nuclei. The total energies in the
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
180.149.51.67 On: Mon, 19 May 2014 08:56:20
224110-6 Lorenz S. Cederbaum J. Chem. Phys. 138, 224110 (2013)
Born-Oppenheimer approximation simply take on the appear-
ance
E
BO,q,,K
=
el
(q + 3/2) + ( + 3/2) +

2
K
2
4M
,
(27d)
= [(2+K)/M]
1/2
.
It is seen that for each electronic state specied by the quan-
tum number q, there is a manifold of rotational-vibrational
states specied by = 0, 1, 2, . . . which arise from the mo-
tion of the nuclei in the potential provided by the electronic
energy (26b). It is also seen that only the center of mass of
the nuclei and not that of the whole system separates from
the electronic and internal nuclear motion. This fact is also
reected in the total energy (27d) which exhibits
2
K
2
/4M as
the translational energy.
We now turn to the exact factorization of the total wave-
function of the system into an electronic and a nuclear wave-
function discussed in Sec. II. The new electronic Hamiltonian
H
el
dening the electronic state is given in (21b) and reads
in the present example
H
el
= H
el
+
2
M
(
n
ln )
n
+

2
n
M
+

2
N
4M
+C
K
. (28a)
Here, H
el
is that in (24). As seen above, in the Born-
Oppenheimer approximation the electronic states
q
depend
in the MK-model only on r R
N
. This is also the case for the

q
. Applying H
el
to
q
, one immediately sees that because of

n
= 0, one arrives at
H
el
=
_
H
el
+

2
N
4M
+C
K
_
, (28b)
which can be further simplied, either by employing the gen-
eral result (21c) or explicitly. Indeed,

2
N
4M
=

2
4M

2
R
2
N
=

2
4M

2
r
2
=
p
2
4M
. (28c)
One immediately sees from (28b) and (24) that in the MK-
model
H
el
=
1
2
p
2
+V
el
(r) +C
K
, (28d)
where
=
2Mm
2M +m
is the reduced mass of the electron. Consequently, and E
el
differ from and E
el
only by the reduced mass and the latter
also by C
K
.
The result reads
E
el,q,K
(| R
1
R
2
|)
=
el
(q + 3/2) +
2+k
4
(R
1
R
2
)
2
+C
K
, (28e)

el
= [2k/]
1/2
.
Using this input data, we are now able to complete the solu-
tion of the problem. The general equation (22a) takes on the
following appearance in the MK-model:
_

2
n
M
+E
el,q,K
(| R
1
R
2
|) +C
K
_

q,
(| R
1
R
2
|)
= E
q,,K

q,
(| R
1
R
2
|), (29a)
which is again an harmonic oscillator problem, readily giving
the following nal result:
E
q,,K
=
el
(q + 3/2) + ( + 3/2) +

2
K
2
2(2M +m)
.
(29b)
Needless to say, = with the above solutions are the ex-
act eigenfunctions of the MK-model and the energies E
q, , K
in (29b) are the corresponding exact energies. We would like
to add that exact results can be similarly obtained for systems
of many light and heavy particles interacting via harmonic
forces. The calculations and the presentation of the results are,
however, rather lengthy.
IV. SUMMARY AND CONCLUSION
In the widely used Born-Oppenheimer approximation, an
electronic Hamiltonian is a priori dened which gives rise
to its electronic eigenfunction. The molecular wavefunction
is then approximated as a product of this electronic wave-
function and a nuclear wavefunction. The latter and the ap-
proximate total energy are found as solutions of a nuclear
Schrdinger equation with the electronic energy serving as
the potential driving the nuclear motion. In this work, we in-
vestigate the nding that the molecular wavefunction can ex-
actly be expressed as a product of an electronic and a nuclear
wavefunction originally discussed by Hunter.
10
In the work of
Hunter, these quantities are constructed a posteriori after the
full solution of the Schrdinger equation has been determined.
In the present work, we determine the electronic and nuclear
wavefunctions entering the product ansatz of the exact total
wavefunction from rst principles. This electronic function is
an eigenfunction of a new rather involved electronic Hamilto-
nian, which depends on nuclear momentum operators. As in
the Born-Oppenheimer approximation, the nuclear wavefunc-
tion is determined by the electronic energy as the underlying
potential. The energy corresponding to this nuclear wavefunc-
tion is, however, the exact total energy of the system.
The electronic Hamiltonian and the working equations
have been derived by two complementary methods. Directly
from the Schrdinger equation for the total wavefunction and
via a variational ansatz. Since the electronic wavefunction de-
pends on the electronic and parametrically also on the nu-
clear coordinates, and is normalized in the electronic space at
each nuclear conguration, the variational ansatz is not trivial.
Interestingly, one does not have to impose a priori this nor-
malization condition for the electronic function. This normal-
ization condition rather follows as a result of the variational
computation.
A simple exactly solvable model is analyzed in the Born-
Oppenheimer approximation and in the exact factorization ap-
proach. The study of the model is presented to didactically
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
180.149.51.67 On: Mon, 19 May 2014 08:56:20
224110-7 Lorenz S. Cederbaum J. Chem. Phys. 138, 224110 (2013)
illuminate the physics behind the approach in a rather trans-
parent case.
Until now it is unclear whether the exact factorization ap-
proach will have practical value in solving complex problems.
The main reason lies in the new electronic Hamiltonian which
is rather involved and of unusual structure. It has already been
argued before that the dependence of the potential energy sur-
face for the nuclear motion on the state in question makes the
factorization approach less practical.
13, 15
We hope that the in-
teresting underlying basic idea of factorization will become
fruitful once the electronic Hamiltonian and its physics are
further studied and better understood.
ACKNOWLEDGMENTS
The author thanks Alexander Kuleff and Ying-Chih Chi-
ang for discussions, and the Deutsche Forschungsgemein-
schaft (DFG) for nancial support. The author also thanks
one of the referees for pointing out the unpublished work in
Ref. (47).
1
M. Born and R. Oppenheimer, Ann. Phys. 389, 457 (1927).
2
M. Born and K. Huang, Dynamical Theory of Crystal Lattices (Oxford Uni-
versity Press, New York, 1954).
3
B. T. Sutcliffe and R. G. Woolley, J. Chem. Phys. 137, 22A544 (2012).
4
H. Kppel, W. Domcke, and L. S. Cederbaum, Adv. Chem. Phys. 57, 59
(1984).
5
D. R. Yarkony, Rev. Mod. Phys. 68, 985 (1996).
6
The Role of Degenerate States in Chemistry, Advances in Chemical Physics
Vol. 124, edited by M. Baer and G. D. Billing (Wiley, New York, 2002).
7
Conical Intersections: Electronic Structure, Dynamics and Spectroscopy,
edited by W. Domcke, D. R. Yarkony, and H. Kppel (World Scientic,
Singapore, 2004).
8
G. A. Worth and L. S. Cederbaum, Annu. Rev. Phys. Chem. 55, 127 (2004).
9
R. T. Pack and G. O. Hirschfelder, J. Chem. Phys. 52, 521 (1970).
10
G. Hunter, Int. J. Quantum Chem. 9, 237 (1975).
11
G. Hunter, Int. J. Quantum Chem. 19, 755 (1981).
12
D. M. Bishop and G. Hunter, Mol. Phys. 30, 1433 (1975).
13
J. Czub and L. Wolniewicz, Mol. Phys. 36, 1301 (1978).
14
P. Cassam-Chenai, Chem. Phys. Lett. 420, 354 (2006).
15
B. Sutcliffe, Theor. Chem. Acc. 127, 121 (2010).
16
M. Moshinsky and C. Kittel, Proc. Natl. Acad. Sci. U.S.A. 60, 1110 (1968).
17
J. Breidbach and L. S. Cederbaum, J. Chem. Phys. 118, 3983 (2003).
18
J. Zanghellini, M. Kitzler, C. Fabian, T. Brabec, and A. Scrinzi, Laser Phys.
13, 1064 (2003).
19
T. Kato and H. Kono, Chem. Phys. Lett. 392, 533 (2004).
20
M. Kitzler et al., Phys. Rev. A 70, 041401 (2004).
21
M. Nest, T. Klamroth, and P. Saalfrank, J. Chem. Phys. 122, 124102 (2005).
22
J. Caillat et al., Phys. Rev. A 71, 012712 (2005).
23
A. I. Kuleff, J. Breidbach, and L. S. Cederbaum, J. Chem. Phys. 123,
044111 (2005).
24
J. Breidbach and L. S. Cederbaum, Phys. Rev. Lett. 94, 033901 (2005).
25
M. Nest, Phys. Rev. A 73, 023613 (2006).
26
S. Klinkusch, T. Klamroth, and P. Saalfrank, Phys. Chem. Chem. Phys. 11,
3875 (2009).
27
A. D. Dutoi, L. S. Cederbaum, M. Wormit, J. H. Starcke, and A. Dreuw,
J. Chem. Phys. 132, 144302 (2010).
28
A. D. Dutoi and L. S. Cederbaum, J. Phys. Chem. Lett. 2, 2300 (2011).
29
Y. Zhang, J. D. Biggs, D. Healion, N. Govind, and S. Mukamel, J. Chem.
Phys. 137, 194306 (2012).
30
M. Drescher, M. Hentschel, R. Kienberger, M. Ulberacker, V. Yakoviev,
A. Scrinzi, T. Westerwalbesioh, U. Kleineberg, U. Heinzmann, and F.
Krausz, Nature (London) 419, 803 (2002).
31
H. Niikura, F. Legare, R. Hasbani, M. Y. Ivanov, D. M. Villeneuve, and
P. B. Corkum, Nature (London) 421, 826 (2003).
32
H. Niikura, D. M. Villeneuve, and P. B. Corkum, Phys. Rev. Lett. 94,
083003 (2005).
33
P. B. Corkum and F. Krausz, Nat. Phys. 3, 381 (2007).
34
O. Smirnova, Y. Mairesse, S. Patchkovskii, N. Dudovich, D. Villeneuve,
P. Corkum, and M. Y. Ivanov, Nature (London) 460, 972 (2009).
35
G. Sansone, T. Pfeifer, K. Simeonidis, and A. I. Kuleff, Chem. Phys. Chem.
13, 661 (2012).
36
L. S. Cederbaum, J. Chem. Phys. 128, 124101 (2008).
37
A. Abedi, N. T. Maitra, and E. K. U. Gross, Phys. Rev. Lett. 105, 123002
(2010).
38
A. Abedi, N. T. Maitra, and E. K. U. Gross, J. Chem. Phys. 137, 22A530
(2012).
39
B. R. Johnson, J. O. Hirschfelder, and K. H. Yang, Rev. Mod. Phys. 55, 109
(1983).
40
P. Schmelcher, L. S. Cederbaum, and U. Kappes, in Conceptual Trends in
Quantum Chemistry, edited by E. S. Kryachko and J. L. Calais (Kluwer
Academic Publishers, The Netherlands, 1994).
41
O. Dippel, P. Schmelcher, and L. S. Cederbaum, Phys. Rev. A 49, 4415
(1994).
42
P. Schmelcher and L. S. Cederbaum, Comments At. Mol. Phys. D2, 123
(2000).
43
V. G. Bezchastnov, P. Schmelcher, and L. S. Cederbaum, Phys. Chem.
Chem. Phys. 5, 4981 (2003).
44
P. Schmelcher, L. S. Cederbaum, and H.-D. Meyer, J. Phys. B 21, L445
(1988).
45
P. Schmelcher, L. S. Cederbaum, and H.-D. Meyer, Phys. Rev. A 38, 6066
(1988).
46
T. Detmer, P. Schmelcher, and L. S. Cederbaum, J. Phys. B 28, 2903
(1995).
47
N. Gidopoulos and E. K. U. Gross, preprint arXiv:cond-mat/0502433.
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
180.149.51.67 On: Mon, 19 May 2014 08:56:20

You might also like